Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

35

Floor Diaphragms in
Timber Buildings
Technical Design Guide issued by Forest and Wood Products Australia
04
01 09
WoodSolutions is an industry initiative designed to provide independent,
Building with Timber
in Bushfire-prone Areas non-proprietary information about timber and wood products to professionals
and companies involved in building design and construction.
BCA Compliant Design and Construction Guide
Technical Design Guide issued by Forest and Wood Products Australia

Timber-framed Construction
for Townhouse Buildings Timbe
Class 1a
Design and construction guide for BCA compliant
Desig
r Floo
n guide ring WoodSolutions is resourced by Forest and Wood Products Australia
for inst
allatio
(FWPA – www.fwpa.com.au). It is a collaborative effort between FWPA
Technica
sound and fire-rated construction l Desi
gn Guid
e issu
ed by n
Forest
and Woo
Technical Design Guide issued by Forest and Wood Products Australia d Prod
ucts
Australia

members and levy payers, supported by industry bodies and technical


associations.
This work is supported by funding provided to FWPA by the
WoodSolutions Technical Design Guides Commonwealth Government.
A growing suite of information, technical and ISBN 978-1-925213-33-1
training resources, the Design Guides have been Acknowledgments
created to support the use of wood in the design
Authors: Daniel Moroder, Prof Stefano Pampanin, Prof Andrew Buchanan
and construction of the built environment.
The research and development forming the foundation of this Design Guide
Each title has been written by experts in the field as well as its preparation and production was proudly made possible by
and is the accumulated result of years of experience the shareholders and financial partners of the Structural Timber Innovation
in working with wood and wood products. Company Ltd.

Some of the popular topics covered by the


Technical Design Guides include:
• Timber-framed construction
• Building with timber in bushfire-prone areas
• Designing for durability
• Timber finishes
First published: 2013
• Stairs, balustrades and handrails Revised: November 2016
• Timber flooring and decking
• Timber windows and doors © 2016 Forest and Wood Products Australia Limited.
• Fire compliance All rights reserved.
• Acoustics These materials are published under the brand WoodSolutions by FWPA.
• Thermal performance This guide has been reviewed and updated for use in Australia by TDA NSW.
IMPORTANT NOTICE
More WoodSolutions Resources
While all care has been taken to ensure the accuracy of the information
The WoodSolutions website provides a contained in this publication, Forest and Wood Products Australia Limited
(FWPA) and WoodSolutions Australia and all persons associated with them as
comprehensive range of resources for architects,
well as any other contributors make no representations or give any warranty
building designers, engineers and other design regarding the use, suitability, validity, accuracy, completeness, currency or
and construction professionals. reliability of the information, including any opinion or advice, contained in
this publication. To the maximum extent permitted by law, FWPA disclaims all
To discover more, please visit warranties of any kind, whether express or implied, including but not limited
www.woodsolutions.com.au to any warranty that the information is up-to-date, complete, true, legally
compliant, accurate, non-misleading or suitable.
The website for wood.
To the maximum extent permitted by law, FWPA excludes all liability in contract,
tort (including negligence), or otherwise for any injury, loss or damage
whatsoever (whether direct, indirect, special or consequential) arising out of
or in connection with use or reliance on this publication (and any information,
opinions or advice therein) and whether caused by any errors, defects,
omissions or misrepresentations in this publication. Individual requirements
may vary from those discussed in this publication and you are advised to check
Downloading of these Technical Design Guides is with State authorities to ensure building compliance as well as make your own
restricted to an Australian market only. Material is only professional assessment of the relevant applicable laws and Standards.
for use within this market. Documents obtained must not
The work is copyright and protected under the terms of the Copyright Act
be circulated outside of Australia. The Structural Timber
1968 (Cwth). All material may be reproduced in whole or in part, provided that
Innovation Company, its shareholders or Forest Wood
it is not sold or used for commercial benefit and its source (Forest and Wood
Products Australia, will not be responsible or liable for any
Products Australia Limited) is acknowledged and the above disclaimer is
use of this information outside of Australia.
included. Reproduction or copying for other purposes, which is strictly reserved
only for the owner or licensee of copyright under the Copyright Act,
is prohibited without the prior written consent of FWPA.
WoodSolutions Australia is a registered business division of Forest and
Wood Products Australia Limited.

Cover image: xxxxxxxxxxxxxxxxxxxxxxxxxxxx


Photographer: xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
Contents

1 Introduction 4

2 Terminology 5

3 Displacement Incompatibilities 7

4 Diaphragm Design for Wind Action 8

5 Load Paths in Diaphragms 9

6 Flexible and Rigid Diaphragms 11

7 Design of Structural Elements in Diaphragms 12

8 Horizontal Deflection of Diaphragms 13

9 Connection Between Single Timber Floor Elements 15

10 Force Transfer Between Diaphragms and Lateral Load-Resisting Systems 16

11 Connection Between Timber Diaphragms and Gravity Frames 17

11.1 Gravity and Shear Forces.............................................................................................................. 17


11.2 Out-of-Plane Rocking..................................................................................................................... 17
11.3 Frame Elongation........................................................................................................................... 18

12 Connections between Timber Diaphragms and Walls 20

12.1 Out-of-Plane Rocking 22

13 Connection Details for Timber Concrete Composite (TCC) Floors 23

14 Design of TCC Floor Diaphragm 25

14.1 Tension Ties.................................................................................................................................... 27


14.2 Compression Struts....................................................................................................................... 27
14.3 Connection of the Collector to Walls.............................................................................................. 28

References 31

#35 • Floor Diaphragms in Timber Buildings Page 3


1
Introduction

Floor diaphragms transfer horizontal forces to the lateral load-resisting system. Horizontal
loads, such as wind, applied to the façade will be transferred as line loads to the edges
of the diaphragms. Horizontal loadings cause inertia forces to develop within the flooring
system that have to be carried to the frames, walls or lift shafts and stairway cores.
Differences in the behaviour of the lateral load-resisting systems (deformed shape, discontinuous
geometry, difference in stiffness) induce additional transfer forces in some diaphragms. Roof and
floor diaphragms also carry gravity loads and they link all vertical structural elements together. It is of
paramount importance that diaphragms maintain their force transferring and linking behaviour before,
during and after any horizontal loadings.

The connection to the lateral load-resisting system can be the weak link in diaphragm design.
As discussed in the next section, this force transfer can be compromised by displacement
incompatibilities typical of jointed-ductile systems.

conn
ll ectio
wa LLRS to
n
conn
ectio
collec n to
tor

chord
conn
ectio
r LLRS to
n
cto
lle
co
open
ing
plate
elem
ent

column
chord
m
ea
m e b ctor
fra olle
c
s
ad
l Lo
o nta
column

riz
Ho

Figure 1.1: Definitions of single diaphragm components.

In this Guide, the horizontal action is considered to be caused by horizontal loads such as wind load.
The first part of the Guide presents the terminology, concept and design of timber diaphragms with
their connections to the lateral load-resisting system (LLRS).
The second part reviews a design example of a timber–concrete diaphragm and its connections to the
LLRS. The diaphragm is subjected to the wind load applied perpendicular to its long side.
For the calculation of timber concrete composite diaphragms, engineers prefer a grillage method with
the use of analysis software. Further information on this grillage method or equivalent concrete truss
method can be found in Strut and Tie Seminar Notes1.
An equivalent truss model is recommended for calculating timber diaphragms. Further information can
be found in An equivalent truss method for the analysis of timber diaphragms2.
New alternative connections and further details including their behaviour in experimental test and also
cost comparison can be found in Design of Floor Diaphragms in Multi-Storey Timber Buildings3, and
Seismic design of floor diaphragms in post-tensioned timber buildings4.

#35 • Floor Diaphragms in Timber Buildings Page 4


2
Terminology

Diaphragms can be made from many different materials – plywood panels, stressed-skin
panels, timber concrete composite floors, cross-laminated timber (CLT), solid floor panels,
structural insulated panels – but their main components can be grouped as follows (see
Figure 1.1):
• plate element
• chords
• collectors/struts
• connections to the lateral load-resisting system.
The simplest method to design diaphragms is the horizontal steel girder analogy, where the web is
made by the plate element and the flanges consist of the chords. The plate element with possible
openings transfers the horizontal shear forces. Several single floor elements may have to be linked
together and forces carried around openings or re-entrant corners. The resultant shear forces have
to be collected and conveyed to the lateral load resisting system via the collectors or struts. The
connection of the collector to the lateral load-resisting system has to be designed properly, as it is an
essential part of the load path into the foundations.
As shown in Figure 2.1 and according to Malone and Rice5, the following terminology is suggested:
• Strut: receives shear from one side only
• Collector: receives shear from both sides
• Chord: perpendicular to the applied load and receives axial tension and compression forces.

Figure 2.1: Irregular floor geometry with typical diaphragm elements.

#35 • Floor Diaphragms in Timber Buildings Page 5


Openings are an unavoidable feature of floor plans, as staircases, lift shafts and channels for services
need to go along the height of a building. The position of these openings has to be chosen carefully,
as they influence the behaviour of a building in multiple ways. First, it influences the load path in the
diaphragm, as the shear forces might have to be carried around it. Bigger openings also increase
the flexibility of diaphragms, which influences the behaviour of the structure and the load distribution
into the lateral load-resisting system. In certain positions, openings also cause a separation of the
diaphragm, as the forces cannot be transferred appropriately and the plane element cannot act as a
unit.
In this situation, often the concept of sub-diaphragms or transfer diaphragms is introduced5,6.
These are portions of the main diaphragm and are used to transfer or anchor higher shear stresses
(from openings, re-entrant corners or concentrated loads) into the remaining diaphragm or the lateral
load-resisting system. Often, closer nail spacing and thicker framing members are adopted, but the
sub-diaphragms are essentially designed as regular diaphragms.

#35 • Floor Diaphragms in Timber Buildings Page 6


3
Displacement
Incompatibilities
Displacement incompatibilities within diaphragms or between diaphragms and the LLRS
can damage structures. Hence, design and detailing are essential to consider displacement
incompatibilities within diaphragms or between diaphragms and the LLRS7. The
displacement incompatibilities normally associated to concrete and steel structures
can also be observed in traditional and innovative timber structures8.
Experimental testing has shown that the flexibilities of timber members and steel fasteners can,
in many cases, accommodate the required displacements without compromising the diaphragm
behaviour.
With careful design, well-designed timber diaphragms can easily undergo horizontal loadings without
any damage. In the case of TCC floors or any floors with concrete topping, some additional detailing
may be necessary as detailed in Section 13.

#35 • Floor Diaphragms in Timber Buildings Page 7


4
Diaphragm Design
for Wind Action
Wind loads are obtained from AS 1170.2. Wind pressures applied to the façade
(perpendicular pressure or wind friction) are then simply transferred to the diaphragms
according to their tributary areas.

#35 • Floor Diaphragms in Timber Buildings Page 8


5
Load Paths in Diaphragms

For regularly shaped floor geometries, i.e. rectangular, without big openings or re-entrant
corners, timber-only diaphragms can be designed by using the girder analogy. This implies
that the shear is taken by the web (diaphragm sheeting) and the bending is taken by the
flanges (diaphragm chords). Even though the girder analogy may not be strictly appropriate
for deep beams with anisotropic materials, tests have shown that flange stresses are smaller
than using that approach, providing a conservative design.9,10 Furthermore, shear stresses
develop uniformly over the web, instead of the parabolic shape found in steel girder webs.
Diaphragms can be designed as simply supported or continuous beams, providing that the span-to-
depth ratio is greater than 2. For aspect ratios smaller than 1, the girder analogy is quite conservative,
as the sheeting and joists contribute substantially in the bending resistance. Considering the high
depth of the diaphragm, the chord forces will be small.11 Different authors provide an upper limit for
the span-to-depth-ratio, as the diaphragm may become too flexible. If floors are running over internal
supports and the different diaphragm parts on each side are connected, they can be analysed as a
continuous beam.
As shown in Figure 5.1, diaphragms with simple plan geometries can be calculated by considering a
girder analogy, where the bending is taken by the chord elements in form of tension or compression
forces and the uniform shear is taken by the plate element.

Figure 5.1: Girder analogy for diaphragms.

#35 • Floor Diaphragms in Timber Buildings Page 9


The tension and compression forces in the chords (T and C) can be calculated as follows, where the
terms are shown in Figure 5.1:

M ql 2
T (5-1)
= −C = =
h 8h
The shear flow along the edges of the diaphragm is:

V ql
v (5-2)
= =
h 2h

where:
q = the lateral load applied to a horizontal diaphragm
l = the span of horizontal diaphragm
h = the width of horizontal diaphragm
V = the shear force applied to a diaphragm
M = the design moment of diaphragm.

The loading of the diaphragms is normally considered as distributed uniformly along the length of
the diaphragm. This is the wind load transferred from the façade, or the inertia forces generated
by the mass of the floor itself. In the case of multi-storey structures, vertical offsets can introduce
concentrated loads. Openings, re-entrant corners, offsets or concentrated horizontal forces will disturb
the shear flow and locally higher stresses might arise.

#35 • Floor Diaphragms in Timber Buildings Page 10


6
Flexible and Rigid Diaphragms

The distribution of forces into the lateral load-resisting system depends on the flexibility of
the diaphragm, i.e. rigid or flexible diaphragms. A diaphragm is considered to be flexible if
its deformation is more than twice the average inter-story drift at that level.
In the case of rigid diaphragms, the transferred forces depend on the stiffness of the diaphragm with
respect to the global stiffness of the lateral load-resisting system. In the likely case that the centres of
stiffness and mass are not coincident, torsional effects have to be taken into account.
To define whether the diaphragm is rigid or flexible, the deflection calculation for both the vertical LLRS
and the diaphragm is required. For flexible diaphragms (Figure 6.1) the load can be determined by
using a tributary area approach.
Depending on the geometry, timber diaphragms often behave somewhere between these two extreme
cases. It remains the designer’s choice of which approach to use: to calculate both to obtain an
envelope; to use a beam on elastic support approach (beam and elastic support have the stiffness
characteristics of the diaphragm and lateral load-resisting system respectively); or to use a finite
element method.
The beam-on-elastic-support approach is valid as long as the stiffness of the diaphragm and the
lateral load-resisting system are similar. The finite element method allows study of the shear stress
distribution over the whole plate element, but requires much more effort in modelling and running the
simulation.

Figure 6.1: Flexible diaphragm.

#35 • Floor Diaphragms in Timber Buildings Page 11


7
Design of Structural Elements
in Diaphragms
The structural elements in a timber diaphragm consist of the plate element, the chords and
the collector/strut beams.
Depending on the chosen setup, the plate element consists either of sheeting panels and framing
members or thicker solid type panels. They all have to be joined to act as a single unit. For a
traditional timber joist floor with panels of particleboard or plywood, the strength of the diaphragm
depends on the amount of nailing and the presence of blocking (connection of the panel edges to
adjacent panels). Blocked diaphragms have a stiffness and strength two to three times higher than
the unblocked equivalent so, for multi-storey timber buildings, such floors should always have solid
blocking along all the sheet edges. The buckling of the sheeting is normally prevented by edge
connections.
The capacity checks necessary to design a diaphragm should be performed in accordance with the
relevant timber code and should include:
• the shear capacity of the nails
• the shear capacity of the panel
• the out-of-plane buckling of the panel.
Nail spacing and sheeting thickness is normally dictated from the stress values at the supports of the
diaphragm. As an alternative, the sheeting panels can be glued or screwed to the framing elements;
this can provide a higher shear capacity and stiffer connections.
Chord, collector and strut beams have to transfer tension and compression forces and need to be
designed accordingly. If they consist of single jointed elements, continuity has to be provided by
adequate ductile connections. Chords should be spliced as far as possible away from the point of
maximum moment. Often these elements also work under gravity loads and internal stresses have to
be combined with the lateral forces. Since the direction of wind forces is arbitrary, chords also act as
collectors – and vice versa – depending on the loading direction.

#35 • Floor Diaphragms in Timber Buildings Page 12


8
Horizontal Deflection of
Diaphragms
There are several reasons why the horizontal stiffness of a diaphragm needs to be
calculated. The vertical elements supporting or attached to a diaphragm need to maintain
their load carrying capacity to guarantee structural integrity, and should not be damaged due
to excessive deformation. Furthermore, the distribution of the in-plane loads to the lateral
load-resisting system is a function of the stiffness of the diaphragm as discussed above.
Finally, the dynamic period of the diaphragm can interfere with the dynamic behaviour of the
structure and cause higher modes effect.
The horizontal deflection of diaphragms is the sum of the following single contributions:

Δ1 = flexural deflection of the diaphragm considering the chords acting as a moment resisting couple
Δ2 = deflection due to shear in the panels
Δ3 = deflection of the diaphragm due to fastener slip
Δ4 = deflection of the diaphragm due chord connection deformation.

Equations for the determination of these contributions are provided in NZS 3603.12

5ql 3
Δ (8-1)
1 =
192EAB 2

ql
(8-2)
Δ 2 =
8GBt

(1+ a)men
(8-3)
Δ 3 =
2

∑δ s x
(8-4)
Δ 4 =
2B
where:
q = lateral load applied to a horizontal diaphragm
l = span of horizontal diaphragm
E = elastic modulus of chord member
A = section area of the chord
B = distance between diaphragm chord member
G = shear modulus of the diaphragm sheathing
t = thickness of the diaphragm sheathing
m = number of the sheathing panels along the length of the edge chord
en = fastener slip resulting from the shear force V
a = aspect ratio of each sheathing panel given in the NZS 360312
x = distance of the splice from the origin
δs = splice slip in the chord.

#35 • Floor Diaphragms in Timber Buildings Page 13


The fastener slip en is calculated for the maximum unit shear force at the support. Since all
diaphragms should be designed as elastic, the slip only depends on the slip modulus and spacing
of the fasteners3. Equation (8-3) can be modified where different fasteners are used for the panel-to-
panel, panel-to-chord and panel-to-collector connections.
Given the variety of wooden construction materials and means of connections, the possible
diaphragm setup and the limited amount of experimental tests, the deflection values only provide
a rough approximation of the diaphragm deformation. The verification of a certain deflection
limit, however, is to be set by the designer with knowledge of the affect deflection will have on the
surrounding structure.
The deflection given above is only applicable to simply supported blocked diaphragms with chord
beams. To account for diaphragm irregularities such as variable loads, openings, re-entrant corners,
changes in diaphragm depth or staggered fastener layouts, these equations can be integrated over
parts of the diaphragm7.
The presence of openings, varying nailing pattern or non-uniform forces should be considered when
calculating the horizontal diaphragm deflection. This can be done by modifying the basic deflection
equation by changing the coefficients of the single contributions according to basic beam theory or by
integrating the equation over segments of the diaphragm. If more precise results are required, finite
element analysis might have to be considered.
The design of floors where the diaphragm action is taken by the concrete topping should be
in accordance with the relevant concrete code. Special provisions regarding displacement
incompatibilities will be given later.

#35 • Floor Diaphragms in Timber Buildings Page 14


9
Connection Between Single
Timber Floor Elements
To connect two single floor panels together, the following alternative connection systems are
suggested (see Figure 9.1):
a) nailing (and gluing) of adjacent panels
b) wooden strip in recess between panels with screws or nails
c) inclined fully threaded screws, or regular screws at 90° between joists
d) nailing (and gluing) of panel to the next joist
e) double inclined screws in shear between solid panels
f) tongue and groove with double inclined fully threaded screws.

Figure 9.1: Connection details between floor elements.


These connections are generally valid and are to be designed to guarantee adequate shear transfer.
Because of the displacement incompatibilities mentioned before, special detailing for floor joints close
to the beam–column joints may be necessary.
Gluing must be considered carefully. A nailed joint or screwed joint with glue will become much
stronger, but also much more brittle and less deformable under extreme loads, so there will be many
cases where glue should not be used.

#35 • Floor Diaphragms in Timber Buildings Page 15


10
Connection Between Single
Timber Floor Elements
Because of the variety of building geometries, lateral load-resisting systems, floor
assemblies and the available types of connection, no unique detail solution can be given.
Key aspects to consider for the connection design are the kind of required force transfer
(horizontal shear only or combined with gravity forces) and the type of diaphragm (timber
only or concrete topping).
While the connections mainly have to transfer the horizontal forces deriving from the diaphragm action,
out-of-plane forces of the lateral load-resisting system have to be considered as well (see Figure 10.1).
These forces can be wind suction at leeward walls, inertia forces on the façade, or dragging forces
from the constraint of vertical elements to move with the rest of the structure under a certain drift.

Figure 10.1: Dragging forces on walls from drift in structure with north–south forces resisted
by frames.

#35 • Floor Diaphragms in Timber Buildings Page 16


11
Connection Between Timber
Diaphragms and Gravity Frames
11.1 Gravity and Shear Forces

In this section, timber-only floors running perpendicular to the lateral and gravity frames are
considered. The floor elements have to transfer vertical gravity forces and horizontal shear forces to
the beam, which acts as a collector or strut. For floors sitting between the beams, gravity loads can
be transferred by a timber corbel, a pocket in the main beam or steel hanger brackets. The horizontal
shear forces can be transferred directly by nailing or screwing the sheeting panels to the beam (see
Figure 11.1).

Figure 11.1: Suggested floor-to-frame connections (floor joists flush with beam): a) floor joist
on corbel; b) floor joist in pocket; c) steel bracket/hanger.
Where the floors sit on the beams, gravity forces are transferred by direct contact. Shear forces can
be transferred by using fully threaded screws at 45° angle or by connecting the sheeting to blocking
elements, which are again joined to the beam by screws or steel plate elements (see Figure 11.2).

Figure 11.2: Suggested diaphragm to frame connections (floor joists on top of beam):
e) floor joist sitting on beam – additional blocking required; f) SIP panel on beam;
g) solid timber floor on beam.

11.2 Out-of-Plane Rocking

Where the horizontal load acts perpendicular to the frame, the whole building will undergo a certain
drift (depending on the lateral load-resisting system in this direction) and hence the frame will have to
rotate out of plane. As indicated in Figure 11.3, it is suggested to leave a construction gap between
the floor elements and the beams (also useful for construction tolerance and variances in ambient
conditions). This will allow the beam to rotate, without damaging the timber floor or the connection to
it. The connection between the beam and the floor also needs to transfer the dragging force to rotate
the frame out of plane.

#35 • Floor Diaphragms in Timber Buildings Page 17


Figure 11.3: Construction gap between a timber floor and supporting beam to allow for
rotation: a) undeformed state; b) deformed state.

11.3 Frame Elongation

The formation of gaps at the beam–column joint produces frame elongation. The diaphragm must
be able to extend. This behaviour has to be allowed for without a brittle tearing of the plate element,
as it would cause permanent damage and compromise the shear transfer. The flexibility of the timber
elements and the low stiffness of the steel connections allow for two simple design solutions for
engineered timber floors:
Solution 1: Concentrated gap (see Figure 11.4, blue details):
As the required deformation in the floor level occurs only at the beam–column joint, a joint between
two adjacent floor panels should be positioned accordingly. This joint needs special detailing, whereas
other panel joints can be designed normally.
For floor setups with sheeting panels and slender joists, only the lower part of the joist should be
connected, so that the joist can bend along its height, but still guarantees shear transfer (see Figure
11.5a). If a different floor setup is used where the joist are too stiff, special steel elements can be
used. These should allow the panels to move apart from each other, but still transfer shear forces (an
example is shown in Figure 11.5b). Appropriate gaps in the floor finishing and the wall linings have to
be provided to allow these deformations to occur.

Figure 11.4: Sample design for a concentrated floor gap (blue) and spread gaps and panel
elongation (red). Solutions for a timber-only engineered floor.

#35 • Floor Diaphragms in Timber Buildings Page 18


Solution 2: Spread floor gaps and panel elongation (see Figure 11.4, red details):
As an alternative to a concentrated gap at each beam location, detailing for uniformly spread gaps
can be used. The required deformation will be accommodated by a number of small panel gap
openings and the elongation of the sheeting panel itself. This implies that the panel is relatively flexible
in the direction perpendicular to the span direction).
Two to three floor elements each side of the interested beam–column joint should be connected to
each other by means of metallic connectors such as nails or screws (like an upper joist connection
shown in Figure 11.5c). The connection needs to guarantee full shear transfer between the elements,
but should be flexible enough to allow for a small displacement. Small gaps will hence open in several
panel joints and the sheeting panels will elongate. The sum of all contributions will make up the
required displacement, as demonstrated in recent testing.
Site gluing to connect floor elements should be avoided, as it results in a stiff and brittle connection
that cannot accommodate the required deformations. Furthermore, the panels close to the beam–
column joint(s) should not be connected to the beam to transfer diaphragm forces, as this would
prevent the development of floor gap openings and panel elongations further away from the area of
interest.
The floor finishing should be chosen to be elastic enough to follow the formation of the spread gaps.

Figure 11.5: a) Lower flange connection; b) connection with thin steel plate; c) upper flange
connection.

#35 • Floor Diaphragms in Timber Buildings Page 19


12
Connections between Timber
Diaphragms and Walls
For wall structures, the diaphragm and gravity forces are transferred via the collector/strut
beam to the lateral load-resisting system (see Figure 12.1). The most appropriate connection
detail to link the collector beam to the walls depends on the span direction of the floor. For
floor elements running parallel to the wall, only horizontal forces have to be transferred –
otherwise gravity forces have to be taken as well.

Figure 12.1: Scheme of a typical diaphragm-to-wall connection.


To allow for the required uplift and rotation of the walls, a discrete connection placed at the centre
of the wall should be used. Ideally, a single dowel-type connector with a vertical slotted hole would
overcome all displacement incompatibilities by still transferring the horizontal forces from the
diaphragm. However, a single dowel is not usually suitable because of the magnitude of the forces,
possible splitting of wood with large diameter dowels and the difficulty of providing slotted holes in
timber.
If only horizontal forces have to be transferred to the wall, connections with steel plates and dowels
placed in slotted holes can be used. The plate itself can be fixed by screws, nails, rivets or bolts to
the timber elements. Where gravity forces also have to be conveyed to the wall, a vertical restraint is
necessary. This solution can be achieved by simply connecting the timber beam and wall together
with dowel-type connectors. While a single big diameter dowel is an attractive solution, little is known
regarding its embedment strength. As an alternative, a ring of closely spaced dowels will approximate
a hinge.

#35 • Floor Diaphragms in Timber Buildings Page 20


Table 12.1 summarises four connection details and their properties (refer also to Figure 12.2).
Table 12.1: Possible wall to collector beam connections.

Connection type Force transfer Displacement incompatibilities Comments


Big pin connection Horizontal shear and Rotation is allowed Uplift is not The embedment
gravity allowed strength and
behaviour of large
diameter dowels is
not well known
Slotted steel plate Horizontal force only Rotation is allowed Uplift is allowed This connection
with rivets allows for all
displacement
incompatibilities. Lots
of steelwork required
Ring of dowels Horizontal shear and Rotation is partially Uplift is not Simple solution,
gravity allowed* allowed the flexibility of the
connection allows for
some rotation
Steel profile with Horizontal force only Rotation is allowed Uplift is allowed Possible problems
slotted holes due to friction

* Given the possibility of using oversized holes in the timber and relatively flexible dowel connection, the
rotation of the wall normally can be accommodated for limited drift ratios.
If uplift of the walls is not allowed by the connection and the collector/strut beam is also attached to
other vertical elements, such as columns (see Figure 12.3a), the beam and the diaphragm will both
need to bend. This can be tolerated if the collector beam is flexible enough (i.e. because of its small
section or a long span to the next vertical restraint). The additional re-entering force resulting from
bending should be considered when designing the wall.

Figure 12.2: Suggested diaphragm-to-wall connection details: a) large diameter dowel


connection (timber–timber); b) dowel connection (steel–steel); c) multiple dowel connection
(timber–timber); and (d) steel angle with slotted holes.

#35 • Floor Diaphragms in Timber Buildings Page 21


For multiple dowel connection (ring of dowels), the additional moment coming from the rotational
restraint should be checked in the beam and in the wall design. Because of the oversized hole in the
timber elements, the compact geometry and the relatively small connection stiffness, this connection
should almost behave as a hinge.

Figure 12.3: a) Single wall; and b) wall with external columns.


Where gravity forces have to be transferred to the wall and the uplift of the collector beam has to be
avoided, a wall configuration with external columns as shown in Figure 12.3b can be used. Under
horizontal loading, the wall would rock but the columns would only follow rotation without any uplift.
In this way, gravity and horizontal forces can be transferred directly to the columns by avoiding any
vertical displacement incompatibility. The connection only has to accommodate the rotation of the
columns. The horizontal force transfer from the columns to the wall and the buckling restrain of the
columns itself must be considered appropriately.

12.1 Out-of-Plane Rocking

As in frame structures, the connections between the floor diaphragm, the collector beam and the wall
itself have to transfer not only the shear flow from diaphragm action, but also the drag force from the
out-of-plane deformation of the wall. This will occur when the horizontal load act perpendicular to the
walls and the whole structure deforms in the out-of-plane direction of the walls. Construction gaps
between the floor and the walls must accommodate the displacement incompatibility.

Figure 12.4: Construction gap between the floor and the wall to allow for rotation:
a) undeformed state; b) deformed state.

#35 • Floor Diaphragms in Timber Buildings Page 22


13
Connection Details for Timber
Concrete Composite (TCC) Floors
As a result of the low tensile strength of concrete, tearing forces due to frame elongation and
bending forces due to uplift and rotation of the walls tend to crack the diaphragm topping.
If these cracks become larger, the force transfer is interrupted and the diaphragm action
compromised13. It is essential to design the diaphragm with its connections accordingly.
For frame structures with TCC floors, the displacement incompatibility required from the beam-
column-gap opening can be accommodated similarly to the concentrated floor gap solution already
described for timber diaphragms. As suggested in Figure 13.1, the concrete should be pre-cracked
along the line of the beam–column joint. Unbonded rebars should be placed across the crack,
designed to deform elastically in case of gap opening and to provide shear transfer via dowel action.

Figure 13.1: Suggested detailing for a TCC floor in a frame system.


The diaphragm has to be tied appropriately to the collector beams. One way to do this is shown in
Figure 13.3. The force transfer from the diaphragm to the beam should be guaranteed in the central
portion of the beams, leaving it unconnected close to the beam–column joints (in the disturbed areas
shown in Figure 13.2). In this way, frame elongation will not compromise the force transfer, which starts
away from the disturbed areas where the displacement incompatibility is attenuated. This is especially
important on external beams and columns, as no concentrated gap opening can be guaranteed.
A different solution to avoid the frame elongation problem on a multi-bay frame consists in connecting
the diaphragm only to one bay and letting the diaphragm slide over the remaining beams. This
solution, however, might result in high shear forces at the connection between the diaphragm and the
beam, and requires proper detailing to allow for the sliding of the diaphragm in respect to all other
elements.

#35 • Floor Diaphragms in Timber Buildings Page 23


Figure 13.2: Shear transfer between the concrete topping and beams.
If the floor gap opening occurs along a collector beam or tie back, care is needed as cracking of the
concrete can compromise the force transfer. Ideally, the pre-crack should be placed away from any
connection to the beams.
Figure 13.3 shows a suggested connection between the concrete topping and the collector or frame
beam. The diaphragm shear is introduced to the beam via notched connections used for the TCC
design (see WoodSolutions Technical Design Guide #30: Timber Concrete Composite Floors). If the
concrete topping is connected to the beam directly, the beam has to be designed as a composite
section. As an alternative, an edge joist from the TCC floor can be connected to the frame beam via
a timber-timber connection. Starter bars are required by the code and have to tie the collector/strut
beam to the diaphragm as well as carrying the shear in case of a crack along the interface.

Figure 13.3: Suggested connection between the concrete topping and timber beams.
In wall structures with TCC floors, the force transfer occurs between the wall and the collector beam
and is unaffected by the presence of the concrete topping. The connection between the diaphragm
topping and the collector/strut beam should be designed as described for frame structures.
If no slotted solution can be adopted for the wall connection, an eventual out of plane bending of
the beam has to be considered. Again, if the span to the next vertical restraint is large enough, the
bending should be accommodated in the concrete topping without excessive cracking. The additional
re-centring force in the wall should be considered, as it might give substantial contribution.
The use of concrete diaphragms in structures that undergo beam elongations cause several
complications and special detailing is required13,14. The suggestions provided in this Guide have not
been fully tested and should be applied with proper engineering judgment.

#35 • Floor Diaphragms in Timber Buildings Page 24


14
Design of TCC Floor Diaphragm

To design the TCC diaphragm, a strut and tie model as per Section 7 “Strut and Tie
Modelling” of the AS 3600 Australian Concrete Code has been adopted. For this design
example, only the design for the wind load applied perpendicularly to the long side of the
building has been carried out. For wind loads perpendicular to the short side of the building,
only a conceptual strut and tie model is shown.
The uniformly distributed loads on the windward and leeward façades (Figure 14.1 and Figure 14.2)
for a wind load have been applied on a four metre grid. The resultant forces are transferred by a
collector beam running on the inner side of the staircase into the post-tensioned walls. The 100 mm
concrete topping is reinforced with a ductile mesh (Ø6.75 mm Grade 500 rebars on 200 mm centres
with a resulting reinforcement area of 179 mm2 per metre width).
Wind loads at ULS:

( )
p = pwindward + pinternal h
(14-1)

1 ( m m )
p = 0.35 kN 2 + 0.21kN 2 3.6m = 2.0 kN

m
load on windward façade (14-2)


2 ( m m )
p = 0.25 kN 2 + 0 kN 2 3.6m = 0.9 kN m load on leeward façade (14-3)

The loads applied to the edge of the diaphragm are carried over compression struts into the collector
beam. Several chord beams along the depth of the diaphragm are taking the tension forces, in this
way the forces can be kept relatively low.

Figure 14.1: Strut and tie model for wind loads on the windward façade.

#35 • Floor Diaphragms in Timber Buildings Page 25


On the leeward façade, the wind loads are first carried over the tension ties into the diaphragm. From
there, compression struts carry the forces into the collector beams.

Figure 14.2: Strut and tie model for wind loads on the leeward façade.
There is no unique strut and tie model for a given geometry and load scenario. A minimum of
experience is required to set up a well-balanced model, as incomplete or not well elaborated models
might lack of equilibrium at the nodes, undergo excessive deformation or require load redistribution
because after concrete cracking.
The conceptual strut and tie model for a wind load perpendicular to short sit of the building is depicted
in Figure 14.3.

Figure 14.3: Strut and tie model for wind load perpendicular to the short side of the building.

#35 • Floor Diaphragms in Timber Buildings Page 26


14.1 Tension Ties

The concrete topping is reinforced by a D500DL72 ductile mesh (reinforcing in both direction made
of Grade 500 reinforcing bars), which also satisfies the minimum reinforcement for crack control for
shrinkage and temperature effects as per Clause 9.4.3 in AS 3600:

( )
= 75% ⋅ 1.75− 2.5σ cp bD ⋅10−3 = 0.75⋅1.75⋅100mm ⋅1000mm = 131mm2
A (14-4)
s,min

where
As,min = minimum reinforcement area
σcp = average intensity of effective pre-stress (0 MPa in this case)
b = width of the diaphragm (taken as 1 m)
D = depth of the concrete topping.

As shown in Figure 14.1 and Figure 14.2, the maximum force in the ties is 12+14.4 = 26.4 kN along
the collector beam running parallel to the wall. Along the collector beam, two additional Ø10 Grade
300 reinforcing bars are placed.

(14-5) 102 ⋅ π
Fnt
= φ st
Af y
= 0.8⋅ ⋅ 300 = 48kN ≥ F * = 26.4kN ∴OK
4
where:
Fnt = nominal tension capacity of steel tie
A = area of the tension reinforcement
Φst = capacity factor for tension struts (0.8).
All other ties have forces of maximum 5.4 kN; therefore, a single leg of the ductile mesh with a Ø6.75
mm Grade 500 rebar provides enough strength to transfer the tension forces.

6.752 ⋅ π
F = φ st Af y = 0.8⋅
(14
nt
mm2 ⋅ 500 N mm2 = 14.3kN ≥ F * = 5.4kN 6)
4
To guarantee the force transfer in the tension ties, the reinforcing bars and the ductile mesh have to be
placed with the required overlapping as provided by the code or the manufacturer.

14.2 Compression Struts

The design strength of a concrete strut, neglecting the reinforcement, is:

(14-7)
Fnc
= φ st β s 0.9 f conc
ʹ Aconc

where
Fnc = nominal compression capacity of concrete strut
f’conc = compressive strength of concrete
Aconc = cross sectional area at one end of the strut, considering the thickness as the depth
of the diaphragm slab (see AS 3600 Australian Concrete Code for more detail)
βs = efficiency factor for concrete struts
Φst = capacity factor for compression struts (0.6).
Considering the maximum force in a strut of only 20kN and a thickness of the slab of 100 mm, all
struts are easily verified. Detailed verifications of the struts and the nodal areas are left to the reader.
The reinforcement plan for the concrete diaphragm is shown in Figure 14.4. Appropriate overlapping
of the reinforcing bars and the mesh has to be guaranteed.

#35 • Floor Diaphragms in Timber Buildings Page 27


Figure 14.4: Reinforcement plan.

14.3 Connection of the Collector to Walls

The force transfer between the diaphragm and the wall can be realized in different ways. Two design
solutions, which are also compatible with the gravity force transfer, are shown. Since the uplift and
rotation of the wall is negligible for this design, no special detailing for the connection between the
floor and the lateral load-resisting system is required.
Solution 1– Direct Connection between Concrete Slab and Wall
Solution 1 consists in a diaphragm force transfer via coach screws fixed directly to the LVL post-
tensioned wall. These are then integrated in the concrete slab when it is cast into place. The floor
elements are sitting on corbels or fixed by steel hangers that are directly connected to the wall. The
design of the latter is not shown here.
The diaphragm force to be transferred into the wall is 46.4 kN (26.4 kN from the ties and 20 kN from
the strut). This force is transferred from the concrete topping to the wall through 6 Ø12 coach screws
(Figure 14.5).
According to clause C4.2 of AS 1720.1:2010, and considering an effective timber thickness of
beff = 2 x tp = 192 mm and embedment strength of f’pj = 17 MPa, the connection capacity is as follows:

t (14-8)
P
≥ 8D = 8⋅12mm = 96mm

⎧b f ʹ D ⎫
(14-9) ⎪ eff pj 2 ⎪ ⎧19.6kN ⎫
Qsk = Qskp = min ⎨ ⎬ = min ⎨ ⎬ = 10.6kN
⎪15 f ʹ D 3 ⎪ ⎩10.6kN ⎭
⎩ pj ⎭
where:
tp = penetration length of fastener
Qsk = characteristic capacity for a laterally loaded single bolt in a joint system
Qskp = system capacity for fasteners loaded perpendicular to the grain
f’pj = characteristic value for bolts bearing perpendicular to grain
The second member in the connection is the concrete slab, which can be considered as stiff; hence,
the factor for side plates (k16) can be taken as 1.2.

N dj ≥ N *
(14-10)
N dj = φ k1k13k16 k17nQsk = 0.8⋅1.14 ⋅1.0 ⋅1.2 ⋅1.0 ⋅ 6 ⋅10.6kN = 69.6kN ≥ N * = 46.4kN

where
Ndj = design capacity for joints under direct load
k1 = 1.14 for wind loads as per Clause 2.4.1.1.
k13 = 1.0 factor for end grain effects
k16 = 1.2 factor for side plates
k17 = 1.0 factor for multiple fastener effect.

#35 • Floor Diaphragms in Timber Buildings Page 28


Figure 14.5: Diaphragm force transfer over coach screws.

The strength of the coach screws embedded in the concrete can be checked according Australian
Code AS 2327.1:2003 Composite structures. Part 1: Simply supported beams.
Solution 2 – Connection between Concrete Slab and Wall via Strut/Collector Beam
An alternative solution consists in a transverse LVL beam running parallel to the wall, fixed with bolts to
it. The diaphragm forces are transferred via notched connections from the concrete topping into the
timber beam (Figure 14.6). For the gravity loads, the floor joists are connected to the transverse beam
by steel hangers. The connection of the beam to the wall is designed for the combination of gravity
and horizontal wind loads.

Figure 14.6: Transverse beam with notches for the diaphragm force transfer.

The factored gravity load is:

p = 1.2G +1.5ψ aQ = 6.7 kN m2


(14-11)

and hence the gravity force on the beam considering a tributary area approach is:

4m
F (14-12)
= p ⋅ A = 6.7 kN m2 ⋅ ⋅ 7m = 94kN
2

#35 • Floor Diaphragms in Timber Buildings Page 29


From the wind load a horizontal force of 46.4 kN has to be transferred into the wall. This leads to a
resultant force and respective angle of:

R* = 46.4 2 + 94 2 = 105 kN
(14-13)

94
ϑ = arctan = 63.7
(14-14)
!

46.4

M16 bolts are used to connect the beam to the wall. A joint group JD3 is assumed for the connection
in LVL elements. The thickness of the connected members is 90 mm for the beam and 225 mm for
the wall. Since the Australian Timber Code AS 1720.1:2010 does not provide the situation of a force
transferred on an angle to the grain in between two members running at 90° to each other, as a
conservative approach the resultant force is applied perpendicularly to the beam.
Considering beff = 2 x 90 mm = 180 mm and f’pj = 17 MPa the connection capacity is:

⎧b f ʹ D ⎫
⎪ eff pj 2 ⎪ ⎧24.5kN ⎫
Q Q min
(14-9)
sk
= skp
= ⎨ ⎬ = min ⎨ ⎬ = 16.3kN
⎪15 f ʹ D 3 ⎪ ⎩16.3kN ⎭
⎩ pj ⎭

A connection with 8 M16 bolts is chosen to transfers the load from the beam into the wall:

N dj ≥ R*
(14-10)
N dj = φ k1k16 k17nQsk = 0.8⋅1.14 ⋅1.0 ⋅1.0 ⋅8⋅16.3kN = 119 ≥ R* = 105kN

A different way to connect the collector beam to the wall could be by using inclined fully threaded
screws; this is, however, not covered here.
To transfer the diaphragm force from the concrete topping into the beam, two notched trapezoidal
connections are provided (Figure 14.7). More information on the design of these connections can be
found in the WoodSolutions Technical Design Guide #30:Timber Concrete Composite Floor Systems.

2N dj = 152.4 > Q* = 64.4kN


(14-15)

2N dj = φ k1k4 k6Qk = 0.8 ×1.14 ×1.0 ×1.0 × 83.5kN = 76.2kN


(14-16)

Qk = 0.95 × 90 − 2 = 83.5kN
(14-17)

where:
Qk = characteristic strength of the TCC connection in shear

Figure 14.7: Trapezoidal notch.


Source: WoodSolutions Technical Design Guide #30:Timber Concrete Composite Floor Systems.

#35 • Floor Diaphragms in Timber Buildings Page 30


References

1. Bull, D.K. and R. Henry, Strut and Tie. Seminar Notes. Technical Report No.R57., in Bulletin of
the New Zealand Society for Earthquake Engineering. 2014, The New Zealand Concrete Society.
Canterbury Earthquake Royal Commission 2012.

2. Moroder, D., et al. An equivalent truss method for the analysis of timber diaphragms. in
Proceedings of the Tenth Pacific Conference on Earthquake Engineering -Building an Earthquake-
Resilient Pacific. 2015. Sydney, Australia.

3. Moroder, D., et al., Design of Floor Diaphragms in Multi-Storey Timber Buildings. International
Network on Timber Engineering Research. Bath, England, 2014.

4. Moroder, D., et al., Seismic design of floor diaphragms in post-tensioned timber buildings, in World
Conference on Timber Engineering. 2014: Quebec City, Canada.

5. Malone, R.T. and R.W. Rice, The Analysis of Irregular Shaped Structures - Diapghragms and Shear
Walls. 2012: McGraw Hill.

6. Diekmann, E.F., Diaphragms and Shear Walls, in Wood Engineering and Construction Handbook,
K.F. Faherty and T.G. Williamson, Editors. 1995, McGraw-Hill: New York. p. 8.1-8.68.

7. Malone, R.T. and R. Rice, The Analysis of Irregular Shaped Structures Diaphragms and Shear
Walls. 2011: McGraw Hill Professional.

8. Vogt, T., J. Hummel, and W. Seim. Timber framed wall elements under cyclic loading. in 12th World
conference on timber engineering, Auckland. 2012.

9. Smith, P.C., D.J. Dowrick, and J.A. Dean, Horizontal timber diaphragms for wind and earthquake
resistance. Bulletin of the New Zealand Society for Earthquake Engineering, 1986. 19(2): p. 135-
142.

10. ATC, Guidelines for the design of horizontal wood diaphragms. Vol. ATC 7. 1981, Berkeley,
California: Applied Technology Council.

11. Prion, H.G.L., Shear Walls and Diaphragms, in Timber engineering, S. Thelandersson and H.J.
Larsen, Editors. 2003, J. Wiley: New York. p. 383-408.

12. Standards New Zealand, Timber Structures, in NZS 3602:1993. 1993, Standards New Zealand:
New Zealand.

13. Bull, D.K., Understanding the complexities of designing diaphragms in buildings for earthquakes.
Bulletin of the New Zealand Society for Earthquake Engineering, 2004. 37(2): p. 70-88.

14. Standards New Zealand, 1170.5 Supp - Structural Design Actions Part 5: Earthquake Actions -
New Zealand - Commentary. 2004b: Wellington, New Zealand.

Australian Standards
AS 1720.1, Timber structures, in Part 1: Design methods. 2010, Standards Australia: Australia.
AS 1170.2:2002 Structural Design Actions Part 2: Wind Actions. 2002, Standards Australia, Australia.
AS 2327.1:2003 Composite structures, Part 1: Simply supported beams 2003, Standards Australia,
Australia.
AS 3600 Concrete structures, 2009, Standards Australia, Australia.

WoodSolutions Design Guide


WoodSolutions Technical Design Guide #30, Timber Concrete Composite Floor System.
WoodSolutions, 2015, Melbourne, Australia.

#35 • Floor Diaphragms in Timber Buildings Page 31


A
Appendix A - Notations

The symbols and letters used in the Guide are listed below:
a aspect ratio of each sheathing panel given in the NZS 3603.12
A section area of the chord
A area of the tension reinforcement
Aconc cross sectional area at one end of the strut, considering the thickness as the depth
of the diaphragm slab as per AS 3600
As,min minimum reinforcement area
b width of the diaphragm (taken as 1 m)
B distance between diaphragm chord member
beff effective width of member in joint assembly
D depth of the concrete topping
en fastener slip resulting from the shear force V
F gravity force on the beam
Fnt nominal tension capacity of steel tie
Fnc nominal compression capacity of concrete strut
F* design action in tension
f’conc compressive strength of concrete
f’pj characteristic value for bolts bearing perpendicular to grain
fy yield strength of steel
G shear modulus of the diaphragm sheathing
h diaphragm width
l length of diaphragm
k1 modification factors for duration of load given in AS 1720.1
k13 nail connector factor for end grain effects given in AS 1720.1
k16 nail connector factor for plywood or metal side plates given in AS 1720.1
k17 nail connector factor for multiple fastener effect given in AS 1720.1
m thickness of the diaphragm sheathing
M design moment of diaphragm
Ndj design capacity for joints under direct load
N* design action for joints under direct load
P wind load at ULS
P1 wind load on windward façade
P2 wind load on leeward façade
p factored gravity load
q uniformly distributed horizontal load
Qk characteristic strength of the TCC connection in shear
Qsk characteristic capacity for a laterally loaded single bolt in a joint system
Qskp system capacity for fasteners loaded perpendicular to the grain

#35 • Floor Diaphragms in Timber Buildings Page 32


R* resultant force
t thickness of the diaphragm sheathing
tp penetration length of fastener
V shear force applied to a diaphragm
v shear flow along the edges of the diaphragm
x distance of the splice from the origin
βs efficiency factor for concrete struts
σcp average intensity of effective pre-stress (0 MPa in this case)
δs splice slip in the chord
Δ1 deflection due to bending
Δ2 deflection due to shear in the panels
Δ3 deflection due to the connection elements
Δ4 deflection of the diaphragm due chord connection deformation
Δdiaphram diaphragm deflection at mid-span
Δinterstory drift LLRS lateral-ing system drift
ϕst capacity factor for tension struts (0.8)
ϑ angle resultant force transferred into the wall

#35 • Floor Diaphragms in Timber Buildings Page 33


Discover more ways to build
your knowledge of wood
If you need technical information or Technical Publications Seminars and Events
inspiration on designing and building with A suite of informative, technical and From one day seminars featuring
wood, you’ll find WoodSolutions has the training guides and handbooks that presentations from leading international
answers. From technical design and support the use of wood in residential and Australian speakers to international
engineering advice to inspiring projects and commercial buildings. tours of landmark wood projects,
and CPD linked activities, WoodSolutions WoodSolutions offer a range of
WoodSolutions Tutorials
has a wide range of resources and professional development activities.
professional seminars. A range of practical and inspirational
topics to educate and inform design and What is WoodSolutions?
www.woodsolutions.com.au construction professionals. These free, Developed by the Australian forest and wood
Your central resource for news about CPD related, presentations can be products industry for design and building
all WoodSolutions activities and access delivered at your workplace at a time professionals, WoodSolutions is a
to more than three thousand pages of that suits you. non-proprietary source of information from
online information and downloadable industry bodies, manufacturers and suppliers.
publications.

You might also like